Вы находитесь на странице: 1из 30

Chemical Physics 262 (2000) 393422

www.elsevier.nl/locate/chemphys

Tunneling splittings in vibrational spectra of non-rigid


molecules
IX. Malonaldehyde and its isotopomers as a test case for fully
coupled multidimensional tunneling dynamics
V.A. Benderskii a,b, E.V. Vetoshkin a, I.S. Irgibaeva a, H.P. Trommsdor b,*
a
b

Institute of Chemical Physics in Chernogolovka, Russian Academy of Sciences, 142432 Moscow Region, Chernogolovka, Russian
Federation
Laboratoire de Spectrometrie Physique, Universit
e Joseph Fourier de Grenoble and CNRS (UMR 5588), BP 87-38402 St. Martin
d'Heres Cedex, France
Received 4 May 2000

Abstract
Twenty one dimensional potential energy surfaces (PES) and the tunneling coordinate dependent kinematic matrices
of malonaldehyde and of several of its isotopomers (D, 13 C) are constructed in the low-energy region (<3000 cm1 )
using quantum-chemical data. Even though the barrier heights calculated with dierent methods dier strongly (from
2.8 to 10.3 kcal mol1 ), all PES become near identical after scaling, tting only one semiclassical parameter c, which
denes the scales of potential, energy and action. The results of dynamical calculations carried out in dimensionless
variables are therefore quite insensitive to the choice of the quantum-chemical method used for the construction of the
Hamiltonian. Choosing the value of c such that the calculated tunneling splitting in the ground state coincides with the
experimental value, the corresponding barrier height is determined as 4.30 kcal mol1 and the vibrational spectrum of
the transition state is obtained. The perturbative instanton approach developed in the previous papers of this series is
used to solve the dynamical problem without reducing the number of degrees of freedom. The role of all 20 transverse
vibrations in proton tunneling is characterized. The tunneling path and globally uniform semiclassical wave functions
are evaluated from the fourth-order HamiltonJacobi equation and the second-order transport equation. Tunneling
splittings in the ground and low-lying excited states are calculated and isotope eects of H/D and 13C/12C substitutions
are predicted. 2000 Published by Elsevier Science B.V.
Keywords: Multidimensional tunneling; Tunneling splitting; Malonaldehyde

1. Introduction
Malonaldehyde (MA, Fig. 1) is one of the simplest molecules with intramolecular hydrogen transfer. The
two equivalent tautomers, OH    O@ and @O    HO, are separated by a moderately high potential
*

Corresponding author. Fax: +33-476514544.


E-mail address: hans-peter.trommsdor@ujf-grenoble.fr (H.P. Trommsdor).

0301-0104/00/$ - see front matter 2000 Published by Elsevier Science B.V.


PII: S 0 3 0 1 - 0 1 0 4 ( 0 0 ) 0 0 3 1 9 - 0

394

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

Fig. 1. Structures of MA. Geometry in the ground state (left) and transition state (right) according to dierent calculations as indicated. The eigenvectors of mode m8 , associated with the tunneling coordinate X in the transition state, are indicated for the optimized
Hamiltonian. The positions of the oxygen atoms are shown on an enlarged scale.

barrier. The ground state tunneling splitting has been derived from microwave and submicrowave spectroscopy [15], a recent direct measurement providing the most precise value of DE0 647046:208  0:019
MHz [6]. This value is about two orders of magnitude larger than the value expected for a 1D barrier [7]
(see also reviews [8,9]). Indeed, it is well established that in MA the reaction path deviates strongly from the
straight line connecting the equilibrium geometries of the two tautomers and involves substantial displacements along small-amplitude, non-tunneling coordinates. These displacements decrease the barrier
and promote tunneling. MA has thus become a natural benchmark system for theoretical studies of
multidimensional tunneling.
In previous papers of this series a new approach to multidimensional tunneling has been developed and
was applied to a few model systems [1017]. In the preceding paper [17], the essential features of this
perturbative instanton approach (PIA) were summarized and applied to H2 O2 . For H2 O2 , accurate
quantum-chemical data are available, and the frequencies of all transverse vibrations are high as compared
to that of the tunneling mode, which is nearly the pure torsion mode. As a consequence, the minimum
energy path (MEP) is only slightly curved and deviates little from the 1D path. In MA, in contrast, the
tunneling mode does not correspond to any of the ground state normal modes and has a high frequency. A
full multidimensional treatment of the tunneling dynamics is crucial in this case, and this molecule was
therefore chosen as second test case for the validity of the PIA. We will show that a quantitative description
of proton transfer in MA and its isotopomers within the PIA does not only reproduce correctly the

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

395

multidimensional eects in the ground vibrational state, but also predicts mode-specic tunneling splittings
in excited states and their variations under H/D and heavy-atom substitutions.
In Section 2, dierent methods of treating multidimensional tunneling are summarized, in particular as
applied to MA. The rest of the paper is organized as follows. The construction of the vibrational Hamiltonian of MA is presented in Section 3. The solution to the dynamical problem is described in Section 4
and in the Appendices AC. Section 5 presents the description of tunneling dynamics, and a concluding
discussion is given in Section 6.
2. Multidimensional tunneling in malonaldehyde
The eect of transverse vibrations on a tunneling path has been discussed within the intrinsic reaction
coordinate approach [18] by Kato et al. [19]. Carrington and Miller have carried out the rst quantitative
study of vibrationally assisted tunneling in MA [20]. Extracting two OH stretching coordinates from the
other degrees of freedom, i.e. from a bath of non-tunneling small-amplitude motions, these authors use the
reactive path Hamiltonian formalism (RPHF) [2123] to take into account the eect of a bath on tunneling
in a 2D heavylightheavy atom system. Using quantum-chemical data, the vibrationally adiabatic potential was constructed [24], and the dependence of transverse frequencies on the tunneling coordinate was
taken into account in calculating the eective width of the reactive channel. However, the adiabatic approximation, on which the RPHF is based, is valid only in the high-frequency limit, when transverse
frequencies are much higher than the frequency of the longitudinal motion in the wells, and a small curvature of the MEP goes along with smooth variations of transverse frequencies [8,22,23]. Clearly this
condition is not fullled for hydrogen transfer in MA, in which the majority of modes has low frequencies.
In agreement with the generalized Fukui theorem [8,18,25], the extreme tunneling trajectory (ETT, path of
least action) starts along the lowest frequency coordinate. This coordinate does not correspond to the
tunneling mode in the system, which cannot be described in the high-frequency limit. The ETT coincides
with the tunneling coordinate only after passage of the high curvature portion of the path, where transverse
frequencies change drastically. In the low-frequency limit, the ETT has the ``corner cutting'' shape and
deviates strongly from the MEP [2628]. In order to take into account these properties, Makri and Miller
[29] have postulated that the straight-line reaction path, introduced in Refs. [30,31], connects intersection
points of caustics, i.e. boundaries between classically acceptable and forbidden regions. A similar shape of
the tunneling path was used later by Shida et al. [32] and Sewel and Thompson [33]. Since this path does not
satisfy the generalized Fukui theorem, Benderskii et al. [25] have introduced non-classical trajectories,
which start from the caustic intersections along the direction of the lowest frequency mode and then deviate
from this direction as higher frequency modes become successively involved in the motion. The so-called
sudden [8,2628,31] and high curvature [23,34] approximations have been developed for the low-frequency
limit, and model 2D problems were studied in order to elucidate the behavior of the ETT in the wide region
of intermediate frequencies [8,32,3537]. Within these models, the total set of (3N 7) transverse vibrations
(in MA with nine atoms their number equals 20) is replaced by one eective transverse vibration which
simulates the eect of vibrationally assisted tunneling. Obviously this simplied model is not suitable to
describe the mechanism of proton transfer correctly.
The above discussion illustrates the absence of unied methods for multidimensional tunneling dynamics. On the one side, the computational cost of a direct solution of the Schr
odinger equation grows
proportionally to 103N 6 and becomes unacceptably high for polyatomics with a number of degrees of
freedom greater than 10. On the other side, the semiclassical WKB method, commonly used for 1D
barrier problems, is extremely dicult to generalize for multidimensional tunneling because of the still
unsolved general problem of matching the semiclassical solutions on the caustics and their continuations
(see, for example, Refs. [38,39]). Only low-dimensional problems can be studied within the WKB

396

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

approximation [4042]. More recently, progress was made using instanton techniques developed by
Polyakov [43] and Coleman [44] in general eld theory (see reviews Refs. [8,45,46]). Model instanton 2D
problems were studied in Refs. [37,4749]. Smedarchina et al. [5055] were the rst to use a multidimensional instanton approach for proton transfer in MA and tropolone. The 3N 6-dimensional potential
energy surfaces (PES) were constructed from quantum-chemical data and the role of each active transverse
mode was evaluated. Along with the eects of transverse displacements and frequency variations, the
mixing of transverse vibrations along the tunneling path was taken into account. In spite of the obvious
advancement in the study of MA by Smedarchina et al. [50], a reevaluation of their results is required for
the following reasons. A simplied version of the instanton approach was used, which makes it almost
equivalent to the RPHF. Totally symmetric and kinematic couplings were assumed to be small, so that the
eect of weakly coupled modes, which are in resonance with tunneling motion, is neglected.
The simplications of the instanton approach, discussed above, are not required within the PIA developed in the previous papers of this series [1017]. A preliminary analysis of proton transfer in MA was
made in Ref. [10], where it was emphasized that the replacement of the ETT by the MEP strongly overestimates the eect of transverse vibrations. An additional reason to recalculate the results of Ref. [50] is
that the PES used in this work is characterized by an adiabatic barrier height of 10.3 kcal mol1 while
modern quantum-chemical calculations [56,57] predict a much lower barrier of 3.94.2 kcal mol1 .

3. Construction of the vibrational Hamiltonian


3.1. Form of the Hamiltonian from symmetry considerations
In the ground and transition states, planar MA belongs to the Cs and C2v symmetry point groups, respectively. The isodynamic group C2
Cs is isomorphic with the point group of the transition state. The
vibrational Hamiltonian is constructed in the reactive coordinates introduced in Ref. [12]. The rectilinear
wide-amplitude coordinate X belongs to the A0 and B2 irreducible representations in Cs and C2v . This
tunneling coordinate is coupled with the set of 20 small-amplitude, transverse vibrations {Yk }. If the irreducible representations of X and Yk are the same in both symmetry groups, linear (L) XYk coupling exists.
If the representations are the same in the Cs ground state only, the lowest order coupling is gated (G). For
coordinates which do not belong to A0 and B2 representations, the coupling is squeezed (Sq). For vibrations
with L and G type coupling, Sq couplings exist also. The set of transverse modes in MA consists of six
vibrations with L type coupling, eight with G coupling and six with Sq coupling. The potential energy
corresponding to three types of coupling is expanded in a power series of reactive coordinates X ; fYk g. If
X-independent anharmonicities of transverse vibrations are neglected, the potential energy is given by the
following expression:

 X

X
X x2 


Ck
akk 2 2
2
k
1 2 X Yk
Ck X
X Yk
Ck 1 X Yk
V X ; fYk g V0 X
2x2k
2
xk
kL
kG
kL;G;Sq
!




X
X
Ck
Ck0
Ck

akk0 X 2 Yk 2 X
Yk 0 2 X
akk0 X Yk 2 X Yk0
xk
xk 0
xk
k;k 0 L
kL;k 0 G
X

akk0 X 2 Yk Yk0
1
k;k 0 G;Sq

The 2D XY terms for L, G, and Sq coupled vibrations and the YY 0 coupling terms are written in the form
discussed in Refs. [10,12,14]. Apart from interactions of the tunneling coordinate with a harmonic bath,

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

397

which are taken into account in the commonly used approximation of non-interacting transverse vibrations, Eq. (1) includes additionally X-dependent YY 0 interactions. This generalization is essential for MA, in
which these interactions mix the normal modes of the transition state along the MEP. The double-well 1D
potential is presented here in a more general form than the one used in previous studies [10,11,5355]:

m
1 X
Vm 1 X 2
2
V0 X
2 m2;3;4
The PES, Eq. (1), is written in a dimensionless form, so that the units of length and of potential energy must
be dened. As shown in Ref. [42], scaling of all distances of zero-point amplitudes:

1=2
h
0
3
dk
mk Xk
to the tunneling distance a0 , used as the unit of length, makes it possible to introduce the semiclassical
parameter, c, as follows:
c

m0 X0 a20
2h

m0 and X0 are the mass and the eigenfrequency in the minimum of the 1D potential. The dimensionless
transverse frequencies are given by:
xk 2

Xk
X0

The factor 2 in Eq. (5) appears because the dimensionless frequency in the minimum of the quartic potential
is chosen to equal 2. The semiclassical parameter is therefore the unit of action, hX0 is the unit of energy,
and the potential energy is measured in units of c hX0 . The PES of MA, dened by Eqs. (1)(5), is characterized by a total of 143 parameters:
(i) Three parameters of the 1D potential m 2; 3; 4
(ii) 14 L and G coupling constants, Ck
(iii) 20 Sq coupling constants, akk
(iv) 106 YY 0 coupling constants, akk 0 .
The kinematic matrix is postulated to be the unit matrix in the transition state. The X-dependent elements
of the kinematic matrix are uniquely determined by the symmetry of the potential couplings [13] and have
the following form:
Gkk 1 gkk X 2
GXX 1 gXX X 2 ;
(

gkk0 X ;
gXk X 2 ; k  L;
GXk
Gkk 0
gkk0 X 2 ;
gXk X ; k  G;

k  L; k 0  G; k 6 k 0
k; k 0  LL; GG; SqSq; k 6 k 0

This matrix contains 141 kinematic constants. The evaluation of these parameters of the Hamiltonian is the
inverse vibrational problem for small-amplitude classical vibrations. The general method used to solve this
problem is given in Sections 2 and 3 of Ref. [14]. The matrices, which give the normal modes of the ground
and transition state in the X ; fYk g-coordinate frame, are derived using the Eckart conditions: no linear or
angular momentum is generated by displacements along the MEP [14], see also Ref. [21].
According to Eq. (1), the dimensionless displacements along L and G vibrations in the ground state with
respect to the transition state equal Ck =x2k . Since these displacements are known in terms of normal

398

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

coordinates, and both transformation matrices are determined as described in Sections 2 and 3 of Ref. [14],
14 L and G couplings constants can be determined from the comparison of the equilibrium geometries of
the ground and transition states. As noted in Ref. [14], these constants are dicult to extract from
the corresponding elements of the force matrices, because they contain also the sums of all YY 0 contributions. In contrast, the YY 0 matrix elements of the force matrices of both the stationary states have the simple
form:
Fkk0 0 x2k dkk0 akk0 ;

Fkk#0 x2k dkk0

This allows to determine YY 0 coupling constants directly.


As shown in Ref. [10], multidimensional tunneling is characterized by the spectral densities of L and G
vibrations:
X
X
2
2
Ck =x2k ;
qG
Ck =xk
8
qL
kL

kG

The value of qG determines a decrease of the adiabatic with respect to 1D barrier:


Vad# V0# 1 qG

while qL characterizes a lengthening of the MEP with respect to 1D path. The action along the ETT between the minima, derived in Appendix A, involves the contributions of L and G vibrations, which increase
and decrease the 1D action, respectively:
X
X
2
2
Ck =x2k F L xk
Ck =xk F G xk
10
W  W0
kL

kG

The functions F xk depend on the shape of the 1D potential and grow with increasing frequency such that
F 0 0, F 1 1. Eq. (10) shows that contributions of transverse vibrations are dened by the spectral
densities, Eq. (8), in the adiabatic limit. When xk is nite, the corresponding contribution decreases. Because of the universality of the F-functions for all transverse vibrations associated with the PES, Eq. (1),
modied spectral densities were introduced in Ref. [10]:
X
X

2
Ck =x2k F L xk ;
q~G
Ck =xk F G xk
11
q~L
kL

kG

Since the PES, Eq. (1), is chosen such that the deformation of the ground state geometry depends only on
the XY coupling constants of L and G vibrations, the spectral densities of both types can be directly found
from a comparison of the ground and transition state equilibrium geometries [14].
The quantum kinetic energy operator is given in the MeyerG
unthard form [58]:


 X

o
1 oG
o
1 oG
o
1 oG
o

GXX X

GXk X
oX 4G oX
oX 4G oX
oX 4G oX
oYk
k
!
X
o2

Gkk 0 X
oYk oYk0
kk 0
2

h
T^
2

12

The kinematic matrix, Eq. (6), is independent of transverse coordinates and GX detG. Through Eq.
(12) the eect of kinematic couplings is reduced to the appearance of additional pseudopotential terms in
Eq. (1) and to the renormalization of potential couplings [13] (see also Appendix B). Since anharmonicities

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

399

of transverse vibrations are not taken into account in Eq. (1), the PES, constructed as described above, is
adequate in the low-energy region, where vibrational quantum numbers equal 0 or 1.
3.2. Evaluation of the Hamiltonian parameters from quantum-chemical data
The equilibrium geometries, eigenfrequencies and energies of the ground and transition states of MA
have been the object of the numerous quantum-chemical calculations [24,32,35,5766]. The eect of electron correlation in the description of the geometrical features is well visualized by the dierence between the
results of HartreeFock (HF) methods and of the MollerPlessett perturbation approach at the MP2 level
[59]. Underestimation of electron correlation in HF methods leads to a longer proton transfer distance,
mainly due to a proton displacement perpendicular to the OO-direction (Fig. 1), and values of the barrier of
10.37.9 kcal mol1 that are higher than the upper limit of 6.1 kcal mol1 determined from proton magnetic
resonance data [67]. The eect of electron correlation is to shift the ground state structure towards the
transition state structure. As a result, the perpendicular proton displacement decreases and the barrier is
lowered. On the other extreme, several recent studies show that standard density functional (DF) calculations overestimate H-bond interactions and predict a distance between proton donor and acceptor sites
that is too short and a XH bond length that is too long [57,68]. On the base of the highest level calculations
with the coupled-cluster method CCSD(T), Barone and Adamo [56] and Sadhukhan [57] recommend best
estimates of the barrier height of 4:3  0:2 and 3.9 kcal mol1 , respectively. These calculations [56,57] show
that the data of the hybrid Becke three parameter LeeYangParr (B3LYP) exchange-correlation functional approach [69,70] are relatively insensitive to an extension of the basis set from 6-311G** up to
6-311G (2df,2p) and give a smaller value of the barrier height of 3.02.8 kcal mol1 as compared to the
more sophisticated and computationally expensive coupled-cluster method.
For these reasons, we choose computationally cheap methods, which give barriers heights close to the
recommended value, and compare the parameters of the Hamiltonians constructed from quantum-chemical
data obtained with dierent methods. The results of this comparison are presented in Tables 1 and 2. The
equilibrium geometry of the hydrogen-bonded fragment in the ground and transition states, calculated with
HF/6-31G**, MP2/6-311G**, MP4SDQ/6-31G** and B3LYP/6-311+G (2dp) methods, conrms the above
mentioned characteristics of the calculations at dierent levels. Although the calculated barrier heights vary
from 10.3 to 2.8 kcal mol1 and the corresponding values of tunneling splitting change by more than three
orders of magnitude, the important parameters of the dimensionless PES, i.e. the spectral densities given by
Eqs. (8) and (11) as well as the dimensionless frequencies, Eq. (5), are almost the same.

Table 1
Ground and transition state geometry of the hydrogen-bonded fragment of MA calculated with dierent quantum-chemical methods
and corresponding barrier heights
Structure parameter


O1 H1 (A)

O2 H1 (A)

O1 O2 (A)
\C1 O1 H1 (deg)
\C2 C1 O1 (deg)
V# (kcal mol1 )
a
b

Ref. [2].
cc-pVTZ basis set [57].

Ground state,
experimenta

Ground//transition state
HF 6-31G**

MP2 6-311G**

MP4SDQ
6-31G**

B3LYP
6-311+G(2d,p)

VSXCb

0.969
1.68
2.574
106.3
124.5
6 6.1

0.956//1.188
1.881//1.188
2.681//2.323
109.40//102.78
126.40//121.98
10.34

0.991//1.198
1.682//1.198
2.585//2.358
104.70//100.56
124.40//121.89
3.45

0.981//1.201
1.780//1.201
2.645//2.356
106.41//101.51
125.20//121.91
5.96

1.002//1.209
1.670//1.209
2.570//2.369
106.07//101.89
123.88//121.65
2.86

0.992//1.215
1.723//1.215
2.607//2.381
106.24//101.53
124.31//121.76
3.70

400

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

Table 2
Characteristics of the PES given by Eq. (1) for MA constructed from data of dierent quantum-chemical methods

Adiabatic barrier height


(kcal mol1 )
Frequency in minima of 1D
potential (cm1 )
Ratio of 1D and adiabatic
barrier heights, Eq. (15)
Spectral density of G-vibrations,
Eq. (14)
Modied spectral density of
G-vibrations, Eq. (17)
Spectral density of L-vibrations,
Eq. (14)
Modied spectral density of
L-vibrations, Eq. (17)
Tunneling splitting in the
ground state (cm1 )
Semiclassical parameter, Eq. (4)

B3LYP 6311+G(2d,p)
this work

MP2 6311G** this


work

MP4SDQ 631G** this


work

HF 6-31G**
this work

VSXC ccpVTZ Ref.


[57]

Scaled
PES

2.86

3.45

5.96

10.34

3.70

4.30

3127

3147

3241

3380

3155a

3025

3.912

3.894

3.705

3.689

3.855a

3.79

0.738

0.743

0.730

0.729

0.739a

0.741

0.317

0.319

0.317

0.311

0.316a

0.317

0.039

0.043

0.032

0.024

0.038a

0.034

0.018

0.019

0.014

0.012

0.017a

0.015

123.2

66.7

5.06

0.05

51.2a

21.6

5.81

6.49

9.56

14.46

6.80a

7.74

Calculated from equilibrium geometries of the ground and transitions states found by Sadhukhan et al. [57], assuming that eigenfrequencies and eigenvectors coincide with those computed with MP4SDQ/6-31G**.

Normal frequencies of the ground and transition states, calculated with these methods, are presented in
Table 3. 1 The average deviation from the experimental data decreases as the predicted barrier height
decreases, but even for B3LYP/6-311+G (2dp) calculations the dierences for the high-frequency active
vibrations remain as large as 200 cm1 . In contrast, the dimensionless transverse frequencies are less sensitive to the method of calculation. When the barrier height ranges from 6.0 to 2.8 kcal mol1 , the variations
of the dimensionless frequencies do not exceed 5% (Table 4), and the description of the vibrations of the
transition state in terms of those of the ground state is nearly the same. This similarity means that dierent
PES derived from the quantum-chemical data, using the dierent methods listed above, are made almost
identical by scaling only one semiclassical parameter. The value of latter can be tted in such a way that the
calculated tunneling splitting in the ground state coincides with the experimental value [16]. The tunneling
splitting in the ground vibrational state is given by the expression [8,10]:
p
13
D0 hX0 cBf0k g exp cW 
In this equation, the prefactor Bf0k g is c independent, and the action, W , between the minima depends on
the modied spectral densities, Eq. (11), and the shape of the 1D barrier. If both these characteristics are
unchanged, ln D0 is a linear function of the semiclassical parameter c as the unit of action. Table 2 shows
that q~L and q~G calculated with dierent methods are almost the same. The insensitivity of 1D barrier shape
as derived from MP2 and MP4 data was tested with the help of additional calculations. Since both above
conditions are fullled, the optimized value of c 7:74 was determined using the calculated dependence of
ln D0 on c and the experimental value of tunneling splitting (Fig. 2). The linear dependence of the adiabatic
barrier height on c (Fig. 2) is used for calculating the interpolated value of Vad . Table 4 indicates that the
1

The numbering of modes is made with frequencies increasing in each of the symmetry representations of the ground state, m1 m15
for A0 and m16 m21 for A00 . For convenience, the modes are arranged in the same order throughout the paper and the quantum numbers
used for these modes are indicated in Table 5.

1379
510
1092
1339
1444
1588
2848
282
873
980
1255
1652
2960
2855
2865
384
768
966
878
252
1028

d(OH)/d(CH4 )//d(OH)/m(OH), m8
Ring def.//ring def., m2
d(CH)//d(CH3 ), m5
d(CH)/m(C@O)//d(CH), m7
d(CH3 )/m(CC)//d(CH3 )/m(CC), m9
m(C@O)/d(OH)//m(C@O)/d(OH), m10
m(CH2 )//m(CH), m12
Ring def.//ring def., m1
Ring def.//ring def., m3
m(CC)//m(CC), m4
d(CH2 )/m(CO)//d(CH2 )/m(CO), m6
m(C@O)/m(C@C)//m(C@O)/m(C@C), m11
m(OH)//m(OH)/d(OH), m15
m(CH)//m(CH), m13
m(CH3 )//m(CH3 ), m14
s(C@C)//s(C@C), m17
c(CH3 )//c(CH), m18
c(CH)//c(CH), m20
c(OH)//c(OH), m19
s(CC)//s(CC), m16
c(CH)//c(CH), m21

1497//1780i
530//614
1211//1201
1535//1427
1600//1629
1770//1717
3178//3294
271//682
957//1030
1035//1140
1378//1499
1928//1812
3938//2053
3369//3293
3395//3421
424//420
857//823
1131//1135
829//1429
248//391
1169//1189

HF

m0 //m#
1440//1258i
513//580
1115//1112
1421//1358
1483//1516
1667//1709
3028//3138
267//635
896//953
1006//1078
1299//1393
1718//1674
3312//1908
3211//3137
3272//3290
376//354
777//771
1004//968
897//1326
273//377
1045//1046

MP2
1454//1493i
507//578
1136//1140
1444//1363
1510//1541
1694//1683
3071//3187
272//637
899//959
1003//1096
1305//1406
1766//1699
3547//1940
3266//3187
3306//3331
378//362
779//768
992//984
856//1334
258//379
1055//1069

MP4
1389//1164i
524//579
1118//1115
1405//1331
1481//1507
1627//1612
2970//3084
277//620
896//950
1002//1059
1287//1370
1691//1633
3214//1897
3107//3084
3174//3230
399//395
784//779
1043//993
1019//1311
288//370
937//1043

B3LYP

1374//1290i
485//550
1068//1068
1359//1292
1420//1450
1595//1613
2894//3001
256//604
852//908
954//1031
1236//1329
1652//1600
3242//1826
3073//3000
3122//3142
358//340
740//731
949//926
836//1263
253//359
997//1003

Scaled

1295i
579
1092
1273
1475
1606
2953
666
930
1059
1349
1601
1667
2788
2883
364
760
943
1327
357
1034

PES-IIc

x#

Ref. [60], see also Ref. [56].


The frequencies are calculated with an interpolation between MP2 and MP4 data using the optimized value of c and a scaling factor 0.95.
c
The vibrational frequencies of the transition state were calculated from experimental data for the ground state and the calculated ratios of scaled frequencies of ground
and transition states.

x0 , experimenta

Normal modes of ground//transition state

Table 3
Normal frequencies (cm1 ) in the ground (m0 ) and transition states (m# ) of MA

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422


401

402

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

Table 4
Dimensionless frequencies of the PES given by Eq. (1) for MA constructed from data of dierent quantum-chemical methods
Normal modes of ground//transition state

x0 //x#
HF

MP2

MP4

Scaled

PES-II

d(OH)/d(CH4 )//d(OH)/m(OH)
Ring def.//ring def.
d(CH)//d(CH3 )
d(CH)/m(C@O)//d(CH)
d(CH3 )/m(CC)//d(CH3 )/m(CC)
m(C@O)/d(OH)//m(C@O)/d(OH)
m(CH2 )//m(CH)
Ring def.//ring def.
Ring def.//ring def.
m(CC)//m(CC)
d(CH2 )/m(CO)//d(CH2 )/m(CO)
m(C@O)/m(C@C)//m(C@O)/m(C@C)
m(OH)//m(OH)/d(OH)
m(CH)//m(CH)
m(CH3 )//m(CH3 )
s(C@C)//s(C@C)
c(CH3 )//c(CH)
c(CH)//c(CH)
c(OH)//c(OH)
s(CC)//s(CC)
c(CH)//c(CH)

0.886//1.053i
0.314//0.363
0.716//0.710
0.908//0.844
0.946//0.964
1.049//1.018
1.880//1.949
0.160//0.404
0.566//0.609
0.612//0.675
0.815//0.887
1.141//1.072
2.330//1.215
1.993//1.949
2.009//2.024
0.251//0.249
0.507//0.487
0.669//0.672
0.491//0.846
0.147//0.231
0.692//0.704

0.915//0.799i
0.326//0.369
0.709//0.707
0.903//0.863
0.942//0.963
1.061//1.088
1.924//1.994
0.170//0.404
0.569//0.606
0.639//0.685
0.826//0.885
1.092//1.064
2.105//1.213
2.041//1.994
2.079//2.091
0.239//0.225
0.494//0.490
0.638//0.615
0.570//0.843
0.173//0.240
0.664//0.665

0.897//0.921i
0.313//0.356
0.701//0.703
0.891//0.841
0.932//0.951
1.047//1.041
1.895//1.967
0.168//0.394
0.555//0.592
0.619//0.676
0.805//0.868
1.090//1.049
2.189//1.200
2.016//1.967
2.040//2.056
0.233//0.244
0.481//0.474
0.612//0.607
0.528//0.823
0.160//0.234
0.651//0.660

0.908//0.853i
0.322//0.365
0.707//0.707
0.900//0.855
0.940//0.960
1.057//1.070
1.916//1.986
0.169//0.399
0.564//0.600
0.631//0.682
0.818//0.878
1.092//1.058
2.143//1.207
2.032//1.984
2.064//2.078
0.237//0.225
0.489//0.483
0.627//0.612
0.552//0.835
0.167//0.237
0.659//0.663

0.912//0.856i
0.338//0.384
0.723//0.723
0.886//0.843
0.956//0.976
1.053//1.065
1.885//1.955
0.186//0.441
0.577//0.615
0.648//0.700
0.830//0.892
1.092//1.058
1.957//1.102
1.888//1.843
1.894//1.906
0.254//0.241
0.508//0.502
0.639//0.623
0.581//0.878
0.166//0.236
0.680//0.684

dimensionless frequencies increase slightly with increasing c. Their values, interpolated between the MP2
and MP4 data, are calculated using the optimized value of c and are scaled by factor 0.95, in agreement
with the recommendation by Chiavassa et al. [71]. The scaled values are listed in Tables 3 and 4. The same
scaling factor is used in order to obtain the optimized value of the adiabatic barrier height as 4.30
kcal mol1 , which is given, together with the other parameters of the scaled PES, in Table 2. When two
conditions indicated above are valid, the values of the adiabatic barrier height and of the characteristic
frequency of 1D barrier are related as follows:
Vad
c
1 qG
hX0 4

14

This relation follows directly from Eqs. (1) and (9). Since all other parameters are known, the value of X0
given in Table 2 is obtained from Eq. (14).
In spite of the frequency scaling described above, the dierence between calculated and experimental
frequencies amounts to 5%. In order to evaluate the eect on the tunneling dynamics of variations, within
this range, of the transverse frequencies, an additional PES-II was reconstructed in such a way that all
normal frequencies in the ground state are taken to be equal to the experimental values using the calculated
ratios x0 =x# . The eigenfrequencies of PES-II are also given in Tables 3 and 4.
Because of the scaling behavior of the PES with respect to quantum-chemical data, we calculate also the
principal parameters of the optimized PES using the VSXC data, which reproduce the CCSD(T) results
[57], in combination with the optimized eigenfrequencies and eigenvectors.
Let us discuss the characteristics of the thus constructed Hamiltonian. 2 In agreement with the nding by
Smedarchina et al. [50], the 1D potential, Eq. (2), is very close to a quartic potential: the ratios of V3 =V2 and
2

The complete set of parameters of this Hamiltonian can be obtained from the authors.

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

403

Fig. 2. Linear relationship between (1) the tunneling splitting in the ground vibrational state, and (2) the adiabatic barrier height on
the value of the semiclassical parameter c calculated from the data of the dierent quantum-chemical methods: (a) B3LYP 6311+G(2d,p), (b) MP2 6-311G**, (c) VSXC cc-pVTZ, (d) MP4SDQ 6-31G**, (e) HF 6-31G**. The experimental value of tunneling
splitting and corresponding values of Vad 4:30 kcal mol1 and c 7:74 are indicated by open squares.

Fig. 3. Proton displacement and deformation of the hydrogen-bonded fragment in Cartesian coordinates, calculated for the optimized
Hamiltonian. (1) 1D trajectory, (2) MEP, (3) ETT. The displacements of the oxygen atoms are shown on an enlarged scale.

V4 =V2 equal 0.067 and 0.046, respectively. It is worthwhile to notice that the rectilinear tunneling coordinate in the X ; fYk g frame does not correspond to a straight-line path in Cartesian coordinates. The
proton displacement and the heavy-atom skeleton deformation along the 1D path, the ETT and the MEP
are shown in Fig. 3. Even the 1D tunneling path is strongly curved due to YY 0 couplings, which produce a
X-dependent mixing of the wide-amplitude mode m8 with transverse vibrations. The 1D barrier height, 16.97
kcal mol1 , exceeds the adiabatic barrier by a factor of about four. This decrease of the adiabatic barrier
results from the strong G couplings.
The L, G, and Sq coupling constants, listed in Table 5, show the following: (i) linear couplings are so
weak that higher order squeezed couplings become dominant for all L vibrations; (ii) strong G coupling
with the lowest frequency mode m1 is partially compensated by the negative Sq coupling; (iii) the contributions that promote tunneling the strongest result from G couplings and positive Sq couplings with m15 ;
(iv) resonance eects occur for m13 and m14 vibrations; (v) all low-frequency Sq vibrations are inactive. The
X-dependent diagonal elements of the kinematic matrix indicate that the eective mass of all motions

404

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

Table 5
Potential L, G, and Sq couplings between tunneling and transverse coordinates in MA
Modes of transition
state

Quantum number

Type of coupling

Dimensionless
frequency xk

L,G-coupling
constant Ck =xk

Squeezed coupling
constant akk =xk

d(OH)/m(OH), m8

2.0

Ring def., m2
d(CH3 ), m5
d(CH), m7
d(CH3 )/m(CC), m9
m(C@O)/d(OH), m10
m(CH), m12

n1
n2
n3
n4
n5
n6

L
L
L
L
L
L

0.365
0.707
0.855
0.960
1.070
1.986

0.020
0.038
0.048
0.074
0.148
0.007

0.108
0.285
0.214
0.138
0.185
0.108

Ring def., m1
Ring def., m3
m(CC), m4
d(CH2 )/m(CO), m6
m(C@O)/m(C@C), m11
m(OH)/d(OH), m15
m(CH), m13
m(CH3 ), m14

n7
n8
n9
n10
n11
n12
n13
n14

G
G
G
G
G
G
G
G

0.399
0.600
0.682
0.878
1.058
1.207
1.984
2.078

0.602
0.210
0.197
0.211
0.062
0.447
0.172
0.137

0.405
0.157
0.048
0.182
0.120
0.421
0.048
0.017

s(C@C), m17
c(CH), m18
c(CH), m20
c(OH), m19
s(C@C), m16
c(CH), m21

n15
n16
n17
n18
n19
n20

Sq
Sq
Sq
Sq
Sq
Sq

0.225
0.483
0.612
0.835
0.237
0.663

0.048
0.024
0.082
0.122
0.077
0.017

changes along the tunneling path. The X-dependence of the non-diagonal elements gkk0 arises from the high
curvature in Cartesian coordinates, mentioned above, of the tunneling path (see Fig. 3). Kinematic couplings are dominant in the interaction between the tunneling coordinate and L vibrations m2 , m9 , and m10 .
Both active G vibrations, m1 and m15 , have strong kinematic couplings and are kinematically coupled with
each other.
As pointed out by Smedarchina et al. [50], mixing of the normal modes of the ground state in the more
symmetric transition state is one of the characteristics of the multidimensional PES of MA. The strong
couplings between the two L vibrations, m7 and m9 , as well as between L vibrations, m7 , m9 , m10 , m12 , and G
vibrations, m6 , m12 , m15 , are ineective due to the weakness of XY couplings for L vibrations. In contrast,
mixing of G vibrations, m1 and m6 , with m8 , m15 , and m4 , m15 , respectively, aects tunneling splittings.
4. Dynamical problem
As shown in Refs. [10,11], the perturbative series converge rapidly when the spectral density is smaller
than 0.5. The high spectral density of G vibrations in MA, qG 0:73, signals that the perturbative solution
of the equations of motion and of the semiclassical equations converges too slowly and that higher than
second-order terms should be taken into account. Moreover, strong potential and kinematic YY 0 couplings
not only mix transverse vibrations but also make them eectively anharmonic. X-dependent anharmonicities of transverse vibrations aect an action and the shape of the ETT. Both eects are expected to give
rise in the semiclassical equations to terms of higher than second-order. For these reasons, the fourth-order
HamiltonJacobi equation (HJE) is used for studying the dynamical problem of proton transfer in MA. A
similar expansion was used in Ref. [17] for the description of tunneling in hydrogen peroxide. For the sake

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

405

of convenience, the perturbative solution of the HJE is sketched in Appendix B. The Nth-order action in the
vicinity of the ETT is expanded over powers of transverse coordinates with X-dependent coecients:

X 2
4
dWk X dWk X   
W X ; fYk g W0 X ; fYk g

X
k

X
kk 0 k 00

1
Fk X



X 1
2
2
Fk X    Yk
Fkk 0 X Fkk0 X    Yk Yk0

kk 0

Fkk0 k00 X Fkk0 k00 X    Yk Yk0 Yk00   

15

The upper index designates the order of HJE and W0 X ; fYk g stands for the action in the separable system
and satises the zero-order HJE:
LR

oW0
oX

2V0 X

1=2

LR

oW0
oYk

xk Yk

16

Two branches of Eq. (16) correspond to the exponents (actions) for the localized wave functions of the left
(L) and right (R) wells [11] (see also Appendix B). The terms of the second sum describe the Y-independent
contributions of transverse coordinates. These terms give rise in the even orders of the perturbative expansion and contribute the action between the minima. The potential and kinematic L and G couplings are
responsible for terms of all orders beginning from N 2, while Sq and YY 0 couplings contribute terms of
N P 4. The third sum denes the ETT as the curve of minimum action. The fourth and fth sums lead to
the X-dependent mixing of transverse vibrations and their X-dependent anharmonicities. The functions,
N
N
Fk X ; Fkk0 X ; . . . are found from the consecutive solution of the HJE equations of orders 1; 2; . . . ; N .
1
2
1
2
The analytical expressions of the rst- and second-order functions, Fk , Fk and Fkk0 , Fkk0 , are given by Eqs.
(B.7)(B.9). As shown in Ref. [17], X-independent anharmonicities are signicant only for active transverse
vibrations. Although this is not taken into account in the present study, much more signicant eects arise
from the XY- and YY 0 -couplings, which, as mentioned above, make transverse vibrations anharmonic,
aecting thus also the shape of the ETT. In principle, the third-order ETT, which includes anharmonic
corrections, can be determined from the equations of motion [17]. However, we prefer to solve the fourthorder HJE to obtain the shape of the ETT together with the action in its vicinity. Numerical calculations of
the fourth-order actions, W L and W R , were performed using analytical expressions of the functions,
N
N
Fk X ; Fkk0 X ; . . . These actions determine the exponents of the semiclassical wave functions of the localized states.
The transport equation (TE) must be solved for the calculation of the prefactors of these wave functions.
The perturbative solution of the TE is described in Refs. [11,13]. The results are summarized in Appendix
C. The numerical evaluation of the prefactors includes the following steps [49]: Zero-order wave functions
are chosen as eigenfunctions of the separable system:
Y
0
17
Wnfnk g Un X Hnk Yk
k

Un X are the eigenfunctions of the 1D potential, Eq. (2), with unit mass, and Hnk Yk are harmonic oscillator functions. The functions Un X and their eigenvalues are computed by numerical diagonalization of
the 1D Hamiltonian matrix. The eects of diagonal and non-diagonal interactions, being both additive
because of the linearity of the TE, can be treated separately. Sq couplings and the X-dependence of eective
0
masses for X- and Yk -motions lead to a renormalization of Anfnk g , given by Eq. (C.3). The X-dependent
0
0
factors Gnfnk g are readily evaluated by integration of Eq. (C.4). The substitution of Anfnk g by A~nfnk g means
that the prefactor for the 1D wave function of X becomes dependent on the quantum numbers of transverse

406

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

vibrations due to Sq couplings and variations of the eective mass. Zero-order wave functions are written in
the semiclassical form:
Wnfnk g A~nfnk g exp cW0 X ; fYk g
0

18

0
where W0 X ; fYk g is the zero-order action. The renormalized prefactors A~nfnk g are used as the basis set for
the expansion of TE solutions over the zero-order states with dierent quantum numbers of longitudinal
and transverse motions:
X
1
0
0
anfnk gn0 fn0k g A~n0 fn0 g
19
A~nfnk g A~nfnk g
k

n0 n0k

For calculating the expansion coecients in Eq. (19), auxiliary functions are introduced:
~ nfn g A~0 exp cW0 X ; fYk g W1 X ; fYk g   
W
k
nfnk g

20

These functions dier from the separable functions, Eq. (18), in that the exponent includes the non-separable action given by Eq. (15). The semiclassical wave functions:
Wnfnk g Anfnk g X ; fYk g exp cW X ; fYk g
are expanded over the basis set of the auxiliary functions, Eq. (20), as follows:
X
~ n0 fn0 g   
~ nfn g
anfnk gn0 fn0k g W
Wnfnk g W
k
k

21

22

n0 fn0k g

As shown in Refs. [11,17], the advantage of the auxiliary functions, Eq. (20), is that the expansion, Eq. (22),
involves c0 -order matrix elements of a new operator, Eq. (C.7), whereas conventional perturbation series
involve c1 -order matrix elements of a perturbation operator summing over all non-separable terms in the
Hamiltonian. As a result, the convergence radius of the expansion, Eq. (22), is much larger than that for a
conventional perturbation expansion. The rst- and second-order prefactors are given by Eqs. (C.10) and
(C.12). The highest order terms, which must be kept in expansions, Eqs. (20) and (22), depend on the values
of the coupling constants and can be easily checked through numerical calculations.
The expansion given by Eq. (22) requires the following comments. The set of quantum numbers,
n; fnk g, species the state of the separable Hamiltonian, while an interaction mixes these states. In the
harmonic approximation, the action dened by Eq. (15) is a quadratic form, which becomes diagonal in the
vicinity of the minima of the PES after transformation of the reactive coordinates into the normal coordinates Q0 . The diagonal form of the action follows directly from Eqs. (B.9), (B.12) and (B.13) at t ! 1.
Therefore, an exponent of the wave function, Eq. (22), is reduced to a product of Gaussians in the normal
coordinates of the ground state. This fundamental property of instanton wave functions of the vibrationless
state n 0; fnk 0g, i.e. its asymptotically smooth transformation into the product of harmonic oscillator functions, has been pointed out by Schmid [72] (see also Refs. [8,11]). All low-lying, well-dened
vibrational states have the same property because variables in the Schr
odinger equation are separated in
normal coordinates Q0 . As a result, we conclude that the following transformation takes place:

Y 0 
23
Hnk Qk
Wnfnk g X ; fYk g X !X0 ! Hn Q00
Yk !Y 0
k

Near a saddle point, an action becomes diagonal in the normal coordinates of the transition state Q# where
the functions, Eq. (22), include the products of harmonic oscillator functions of transverse coordinates:

 Y #
24
Hnk Qk
Wnfnk g X ; fYk g X !X # ! N Q#
0
Yk !Y

#
k

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

407


#
N Q#
0 is the wave function of the separable motion along the longitudinal coordinate Q0 . Both products of
the transverse harmonic oscillator functions are characterized by the same set of quantum numbers {nk }.
Transformations, Eqs. (23) and (24), are automatically fullled in the RPHF, where quantum numbers are
postulated to be the same along the MEP, while variations of transverse frequencies and eigenvectors result
from changes of the composition of normal mode due to the wide-amplitude dependent couplings. A
conservation of quantum numbers in the high-frequency limit follows from a decrease of the transition
amplitudes in expansion, Eq. (22), with increasing energy denominators in Eq. (C.10). However, for nite
values of transverse frequencies, when transition amplitudes are suciently large and the ETT deviates
from the MEP, quantum numbers are not conserved. Nevertheless, it is possible to characterize mixed
0
states by the same sets of vibrational quantum numbers if Anfnk g remains the leading term in the expansion,
Eq. (22). In principle, the expansion, Eq. (22), is made over the total set of vibrational states including
highly excited states. L and G couplings induce transitions between states n; fnk g and n0 ; fnk  1g. Since
tunneling splittings increase exponentially with increasing longitudinal quantum numbers, even small
transition amplitudes between 01k and n0 0k states, a0f1k gn0 f0k g , increase appreciably the tunneling splitting
of the rst excited state of the kth transverse vibration. This enhancement grows abruptly in regions of
Fermi-resonances between longitudinal and transverse vibrations [11]. Sq coupling mixes states n; fnk g
and n0 ; fnk  2g, while YY 0 couplings are responsible for the ``double transitions'' n; fnk ; nk0 g $
n0 ; fnk  1; nk0  1g. All transitions, mentioned above, involve excited states of the wide amplitude coordinate X with n 1; 2; . . . States with n P 2 are situated above the top of the barrier. Although the
contributions of these highly excited states can be treated within the PIA [16,17], in the case of MA, their
incorporation in the dynamical problem is questionable, because the PES, Eq. (1), becomes inadequate in
this energy region. However, the transition amplitudes between strongly localized low-lying states and
delocalized X-states with n P 2 are exponentially small and decrease rapidly with increasing longitudinal
quantum numbers. Moreover, the high characteristic frequency of the longitudinal mode makes these
amplitudes much smaller than in hydrogen peroxide, where the tunneling coordinate is associated with the
low-frequency torsion mode. These reasons justify the use of a restricted basis set for studying the proton
transfer dynamics in MA. Nevertheless, we included the state 2f0k g in order to take into account the
mixing of this state with excited states of transverse coordinates and found indeed that, for the reasons
mentioned above, its eect is very small.
0
The basis set of the separable functions, Eq. (17), is orthonormal. However, substitution of Anfnk g by
0
A~nfnk g mixes states with n0 n  2. The overlap integrals hWn0 jWn i for n 0; 1 are exponentially small, so
that the non-orthogonality of the wave functions, given by Eq. (18), can be neglected. Therefore, the basis
set of the orthonormal separable functions includes three functions of the lowest states of the 1D potential,
Eq. (2), and two harmonic oscillator functions, nk 0; 1, for each transverse vibration. The dierent
products of these functions are divided in a prefactor and an exponent using zero-order actions calculated
through Eq. (16). The basis set of the auxiliary functions, Eq. (20), is constructed taking into account the
n; fnk g-dependent renormalization of the prefactors given by Eq. (C.3) and the rst- and second-order
actions calculated with Eqs. (B.7)(B.9) and (B.11)(B.13). The third- and fourth-order actions are also
included for active G vibrations. Even for these strongly coupled vibrations, contributions of fourth-order
terms to tunneling splittings are found to be negligible. The rst-order prefactors for the ground and rst
excited states of all transverse vibrations are calculated using expansion, Eq. (22), with the rst- and second-order coecients, Eqs. (C.10) and (C.12). As a result, 42 globally uniform semiclassical wave functions
of all low-lying vibrational states of MA are computed. Since the transformation matrix of the reactive
coordinates X ; fYk g into Cartesian, mass-weighted coordinates is known, these functions can be represented in a form, which is suitable for calculating transition amplitudes of dierent type intramolecular
transitions. Contour plots of the wave functions for the ground and rst excited states of vibrations m1 and
m15 are shown in Fig. 4. In agreement with Eq. (23), these functions are almost separable near the minima.
As the displacement along X grows, the equilibrium positions for both transverse vibrations are displaced

408

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

Fig. 4. Contour maps of the wave function density on planes spanned by coordinates, X ; Y1 ; Y15 as indicated, of the vibrational levels:
(a) (0,01 ,015 ); (b) (0,01 ,115 ) and (0,11 ,015 ).

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

409

and the vibrations themselves become anharmonic. Near the minimum, the states 0; 11 ; 015 and 0; 01 ; 115
are actually the excited states of the normal vibrations m1 and m15 , but on the dividing plane these states are
rather related to the 1; 01 ; 015 state due to the XY interaction. YY 0 coupling also aects this mixing.
Tunneling splittings are calculated using the generalized LifshitzHerring formula [11], given by the
expression:
0
1
Z
X

hX0
hX0 @ 0
0
Dnfnk g
Wnfnk g rWnfnk g X 0 dfYk g
anfnk gn0 fn0k g Dn0 fn0 g A
25
Dnfnk g
k
c dividing
c
n0 fn0 g
plane
k

Dnfnk g are tunneling splittings in the ``pure'' states corresponding to the eigenfunctions dened by Eqs. (19)
and (20). Since the dierence in matrix elements for odd and even states is neglected, Eq. (25) is valid when
tunneling splittings are much smaller than energy spacings in vibrational progressions. The more rigorous
expression, given in Ref. [17], is not required for MA. In the harmonic approximation, the action given by
Eq. (15) is reduced to a quadratic form, so that integration over the dividing plane in the LifshitzHerring
formula involves a diagonalization of this form. Its eigenvectors do not coincide with those of the ground or
transition states, because the X-dependent coecients depend on the shape of tunneling trajectory. The
variation of a transverse frequency along the ETT due to L, G, and Sq coupling is given by the expression:



gkk 
akk
0
1
2
0
xk xk 1
1 2
26
xk X xk Fkk Fkk ;
2
xk
0

xk is the transverse frequency in the ground state, and Fkk and Fkk are given by Eqs. (B.9) and (B.13),
0
respectively. Since the normalization factors of localized wave functions depend on xk , tunneling splittings
contain transverse-mode-dependent factors, which are independent of action. Thus, both eects, the
variations of action and the X-dependence of the width of the reactive channel, are separated and tunneling
splittings can be written in the form [8]:
!n 1=2
Y x k 0 k
0
0
Bfnk g
27
Dnfnk g Bfnk g Dnfnk g ;
0
xk
k
0

with xk 0 dened by Eq. (26), and where Dnfnk g includes only variations of the action along the ETT. When
YY 0 couplings are taken into account, transverse-mode-dependent factors are no longer separable. The
partition of eects, as described above, is useful in order to clarify the role of dierent types of interactions
in tunneling dynamics.
5. Tunneling dynamics
According to Eq. (B.12), each transverse vibration contributes to the second-order action due to potential and kinematic XY couplings. Both of these contributions are given in Table 6. The eect of kinematic couplings alone is small for all vibrations and only the cross-terms, proportional to gXk Ck , are
signicant in the action. L couplings are small, so that Sq couplings dominate transverse-mode-dependent
factors for L vibrations. Although these factors dier signicantly from unity, the total eect is small,
because Sq couplings of dierent signs partially compensate each other. The largest increase of tunneling
splitting results from vibrations m1 and m15 . These two active vibrations alone increase the 1D splitting in the
ground state by a factor of 8.4, while the combined enhancement resulting from all other modes (except m1
and m15 ) gives a factor of 1.74. For other G vibrations, G and Sq coupling contributions are comparable.
The combined eect of Sq vibrations provides an overall transverse-mode-dependent factor of 0.884.

410

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

Table 6
Contributions of transverse vibrations to tunneling splittings in the ground vibrational state of MA
Ia

IIb

m2 (n1 )
m5 (n2 )
m7 (n3 )
m9 (n4 )
m10 (n5 )
m12 (n6 )
m1 (n7 )
m3 (n8 )
m4 (n9 )
m6 (n10 )
m11 (n11 )
m15 (n12 )
m13 (n13 )
m14 (n14 )
m17
m18
m20
m19
m16
m21

(n15 )
(n16 )
(n17 )
(n18 )
(n19 )
(n20 )

IIIc

IVd

Ve

VIf

0.0008
0.0011
0.0013
0.0027
0.0089
0.0

0.0008
0.0011
0.0014
0.0027
0.0091
0.0

0.0001
0.0004
0.0026
0.0012
0.0045
0.0

0.999
0.997
0.980
0.991
0.966
1.0

0.962
1.164
1.134
0.918
0.887
0.918

0.1024
0.0160
0.0151
0.0197
0.0018
0.1007
0.0171
0.0110

0.1254
0.0165
0.0154
0.0201
0.0019
0.1092
0.0173
0.0111

0.1541
0.0226
0.0122
0.0261
0.0026
0.0983
0.0179
0.0108

3.296
1.191
1.099
1.223
1.020
2.140
1.149
1.087

0.891
0.926
1.025
1.115
1.081
1.332
0.963
0.986

1.012
1.010
0.960
0.931
0.980
0.991

Transverse vibrations in the transition state.


Contribution of second-order potential coupling L(G) to the action along ETT between minima from Eq. (A.9).
c
Contribution of fourth-order potential coupling L(G) to the action along ETT between minima.
d
Total of contributions of potential (column II and III) and kinematic couplings L(G) to the action along ETT between minima.
e
expcdW with W from column IV and c 7:74.
0
f
Sq-factor for ground state, D0 =D0 .
b

Thus, the general scenario of proton transfer in MA can be described as follows. The tunneling coordinate, which is associated with mode m8 in the transition state, stretches the OH bond and decreases COH
angle (Fig. 3). The two promoting vibrations, m1 and m15 , provide the low-frequency ring deformation and
the high-frequency symmetric deformation of the hydrogen-bonded fragment. Eigenvectors of both promoting modes are shown in Fig. 5. These three transition state modes change strongly along the tunneling
path, so that the majority of A0 vibrations of the ground state are involved in the tunneling transition. From
this point of view, it is impossible to separate the active modes from the bath of weakly coupled vibrations.
This mixing leads also to strong variations of the transverse frequencies along the ETT. For active
transverse vibrations, these variations are shown in Fig. 6. It is worthwhile to notice that similar changes,
taking place for the inactive modes, do not reveal themselves in the tunneling dynamics. The X-dependent
tunneling-induced anharmonicities, though quite signicant, also play a minor role in the proton transfer.
As an example, the quartic anharmonicities of active vibrations are indicated in Fig. 7.
The frequencies of the promoting vibrations do not correspond to the adiabatic limit, so that the ETT
deviates from the MEP. The shapes of the ETT and the MEP in 3D congurational space, X ; Y1 ; Y15 , are
compared in Fig. 8. In contrast to the MEP, the ETT starts along the lowest frequency mode m1 and then
turns along the m15 mode. The ETT crosses the dividing plane in a point that is not the saddle point. The
barrier along the ETT is intermediate between 1D and adiabatic barriers (8.09 vs 16.18 and 4.30 kcal mol1 ,
respectively). The decrease of the 1D barrier along the ETT is related to the modied spectral density, Eq.
(11), while the spectral density, Eq. (8), denes the ratio of adiabatic and 1D barriers. The dierence between spectral densities, Eqs. (8) and (10), characterizes the dierence of the second-order action calculated

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

411

Fig. 5. Eigenvectors of active G vibrations, m1 and m15 , in the ground and transition states of MA.

Fig. 6. Variation of transverse frequencies of the active G vibrations, m1 and m15 , along the MEP () and along the ETT (- - -),
calculated with the optimized Hamiltonian.

412

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

Fig. 7. Tunneling induced quartic anharmonicities of vibrations m1 and m15 along the MEP () and along the ETT (- - -), calculated
with the optimized Hamiltonian.

Fig. 8. Plot of the ETT and of the MEP in 3D space X ; Y1 ; Y15 . Transverse displacements are shown on an enlarged scale.

along the ETT and within the adiabatic limit. The modied spectral density for proton transfer in MA is
about two times smaller than that predicted within the adiabatic limit (see Table 2). The tunneling splitting
in the ground state calculated for the 1D barrier, Eq. (2), equals 0.87 cm1 . Therefore, the overall promoting eect of transverse vibrations increases the tunneling splitting by a factor of 25, whereas the value
predicted within the adiabatic limit, 55.73 cm1 , is 2.6 times larger than the experimental value.
Tunneling splittings in all rst excited vibrational levels of MA are given in Table 7. The splitting increases drastically in the rst excited level of the wide-amplitude vibration, because this level is located close
to the top of the barrier. The corresponding splitting of the 1D potential (determined by the diagonalization
of 1D Hamiltonian matrix) is 615 cm1 and decreases in the multidimensional PES to 503 cm1 due to the
coupling with transverse vibrations. Since the exponent of the instanton wave functions, Eq. (21), is independent of quantum numbers, tunneling splittings in excited states of transverse vibrations change due to
variations of the prefactors. The transverse-mode-dependent prefactors, given by Eq. (27), depend on nk

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

413

Table 7
Tunneling splittings (cm1 ) in MA, MA-d1 , MA-d4 , MA-13 C
State

MA

MA (PES-II)

MA-d1

MA-d4

MA-13 C

0
1m8

21.6
503.19

21.8
347.4

3.0
346.14

2.84
202.36

20.60
544.53

1m2
1m5
1m7
1m9
1m10
1m12

19.75
25.57
27.23
23.70
17.27
20.77

19.82
25.90
27.15
23.72
17.25
20.75

2.47
4.27
3.77
3.06
4.29
2.70

3.46
4.50
3.74
4.00
3.79
3.18

18.53
23.77
25.87
22.40
16.31
19.62

1m1
1m3
1m4
1m6
1m11
1m15
1m13
1m14

33.75
26.52
29.71
38.11
22.20
64.02
163.18
127.12

37.37
26.90
30.42
39.04
22.20
52.28
254.21
190.87

6.52
4.00
4.28
6.68
3.20
7.10
6.74
5.30

5.79
4.17
3.19
3.93
6.18
8.15
73.12
60.43

30.05
26.76
29.51
34.24
20.97
55.92
146.22
118.03

1m17
1m18
1m20
1m19
1m16
1m21

22.70
22.47
18.56
15.82
18.03
20.63

22.71
22.47
18.61
15.90
18.02
20.64

3.18
3.15
2.11
3.10
2.52
2.87

3.53
2.91
2.09
1.81
2.47
3.03

21.39
21.19
17.54
14.18
17.50
19.45

and their changes are due to Sq couplings. Tunneling splittings increase (decrease) if the transverse frequency, averaged over the ETT, is higher (lower) than in the well. XY and YY 0 couplings are responsible for
the mixing of ``pure'' states with dierent vibrational quantum numbers. Mixing between states 0; 1k with
small tunneling splittings and state 1; 0k with a large splitting always increase the tunneling splitting in
mixed excited states of L and G vibration. YY 0 couplings average tunneling splittings over the set of excited
levels of transverse modes.
As a result of the dierent factors mentioned above, increases or decreases of tunneling splittings are
predicted in excited vibrational levels of MA, as discussed below. For L vibrations, m2 , m5 , m10 , and m12 ,
mixing is small and tunneling splittings vary according to the sign and the magnitude of the Sq coupling
constants. For the L vibration m9 , due to YY 0 coupling with G vibration m6 , the splitting increases in spite of
a negative Sq coupling constant. Coupling between m6 and m7 increases the splitting in the 0; 17 state. For
G vibration m1 , XY coupling is dominant. As shown in Fig. 4, level 0; 11 is strongly mixed with level 1; 01
even though the energy dierence between these states is relatively large. The resulting increase of tunneling
splitting in level 0; 11 exceeds the eect of the strong negative Sq coupling. For G vibration m3 , the splitting
increases due to XY coupling and YY 0 coupling with vibration m1 . YY 0 coupling between m4 and m6 promotes
tunneling in level 0; 14 . The large splitting in levels 0; 16 and 0; 115 result from a combination of relatively strong XY and positive Sq couplings. Resonance eects on tunneling splittings arise for G vibrations
m13 , m14 , and m15 . The rst two vibrations are very close in frequency to m8 , so that the matrix elements
between levels 1; 013 ; 014 and levels 0; 113 , 0; 114 are comparable to the energy denominators. Due to the
change of transverse frequencies, tunneling splittings vary signicantly when the optimized PES is replaced
by the PES-II (see Table 7). At the level of accuracy at which the PES was constructed in this work, only the
existence of resonances can be predicted, but more reliable values of frequencies are required in order to
quantitatively predict the value of the splittings.

414

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

The features of tunneling dynamics, described above, are also manifest in H/D substituted MA. The
most spectacular change concerns G vibrations m13 and m14 , which are o resonance in MA-d1 and in
resonance in MA-h4 and MA-d4 . As a result, tunneling splittings in states (0113 ) and (0114 ) of MA-d1 are
predicted to be smaller by more than one order of magnitude as compared to the other isotopomers.
6. Discussion
The multidimensional tunneling dynamics of MA, as described in Section 5, is in qualitative agreement
with results of previous studies. The dominant role of G vibrations m1 and m15 for vibrationally promoted
proton transfer in MA has already been indicated by Smedarchina et al. [50]. However, because of the high
value of the barrier used in their calculations (10.3 and 39.3 kcal mol1 for the adiabatic and 1D barriers,
respectively), the eect of G vibrations was strongly overestimated, in order to make the high barrier
suciently transparent and to reproduce the experimental tunneling splittings. This overestimation arises
when the action along the ETT is replaced by the action along the MEP and the spectral density, Eq. (8), is
used instead of the modied spectral density, Eq. (11). The parameters of the model 2D Hamiltonian
studied in Refs. [32,34,36] reproduce correctly the height of the real barrier, 4.98 and 18.01 kcal mol1 for
adiabatic and 1D barriers, respectively. The frequency of the eective transverse vibration x 0:71 was
chosen to be intermediate between m1 and m15 , so that tunneling path evaluated within this simplied model
accounted for the real deviations of the ETT from the MEP. Nevertheless, the actual multidimensional
dynamics is much more complicated than that predicted in Refs. [32,34,36,50]. First, the eect of Xdependent interactions between transverse vibrations, which was not taken into account in earlier work, is
important. Because of YY 0 couplings, contributions of dierent transverse vibrations are strongly mixed not
only in the ground vibrational state but also in the rst excited states of normal modes. Resonances in
tunneling splittings have also been beyond the scope of all previous studies.
As for the dynamics of hindered rotation in hydrogen peroxide, described in the previous paper [17], the
PIA oers also a satisfactory solution of the multidimensional dynamical problem in MA. The validity of
the results is limited by the precision with which the PES can be constructed from quantum-chemical data
rather than the accuracy of dynamical calculation itself. The available ab initio quantum-chemical data do
not permit the construction of a more reliable vibrational Hamiltonian for MA, because the precision with
which barriers and frequencies of normal modes are calculated with dierent methods signicantly aects
the results of the dynamical calculations. Although the scaling behavior of the PES, discussed earlier, makes
it possible to construct an optimized Hamiltonian using the experimental values of normal frequencies and
tunneling splitting, an experimental check of vibrational-tunneling spectra, predicted in this paper, is required in order to prove the correctness of the results. Nevertheless, the two contributions to the tunneling
dynamics of proton transfer in MA pointed out in this paper, namely the existence of resonances and
tunneling induced anharmonicities of transverse vibrations are independent of any variation of the
Hamiltonian parameters. Indeed, the width of the resonances is determined by the values of coupling
constants (Ck =xk and akk 0 =xk xk0 for XY and YY 0 resonances, respectively [11]) and exceeds the dierences in
energy denominators in Eqs. (C.10) and (C.12), calculated with dierent methods.
Below, the role of these two eects in the tunneling dynamics is discussed in more detail. Resonances
increase the frequency of tunneling oscillations between wells in excited levels of transverse vibrations. This
resonant enhancement of coherent tunneling has been described earlier in Ref. [15] for nitric acid, in which
the second excited level of the wide-amplitude torsion mode is in resonance with the rst excited level of the
ONO bending mode. As a result, the tunneling splitting in this level increases by about three orders of
magnitude as compared to the ground state. In MA, a resonance occurs between the antisymmetric
stretching/bending OH vibration m8 and its symmetric counterpart m15 . Two other resonances appear between m15 and the CH-stretch vibrations m13 and m14 . In MA-d1 , the rst resonance prevails, while two others

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

415

cease to exist. All three resonances exist also in MA-d4 . This resonant enhancement of tunneling transitions
in excited levels of transverse vibrations is the coherent analog of mode-specic tunneling reactions. As
noted in Refs. [14,15], this type of mode selectivity oers the opportunity to induce chemical processes with
large, heavy-atom isotope eects. The results for MA-13 C conrm this statement. The isotopic substitution
aects tunneling splittings of excited vibrational states much more than the frequencies of the corresponding modes.
The existence of large YY 0 coupling implies that the approximation of non-interacting transverse vibrations becomes invalid in MA. Tunneling transitions are accompanied by a redistribution of vibrational
energy. A unied method for the description of the time evolution of the initially prepared non-stationary
states is the use of an evolution operator (propagator). Typically, the construction of the semiclassical
propagator for multidimensional deep tunneling problems is very dicult (see [7375] and references cited
therein) due to the same limitations as those mentioned in Section 2 for the WKB method. However, these
limitations are circumvented within the PIA. Once the instanton wave functions, Eq. (21), are computed,
the real-time propagator can be found using its denition (see, for example [76,77]):


X
iEn t
Wn QWn Q0 exp
28
K Q; t; Q0
h

n
where Q X ; fYk g, n n; fnk g. Since highly excited states of longitudinal vibrations are uncoupled
from localized states, the truncated propagator, in which only low-lying states are included, is adequate for
the description of tunneling dynamics. If transverse vibrations are non-interacting, the propagator, Eq.
(28), consists of two terms tracing the fast intrawell oscillations on the time scale of X1
0 and slow tunneling
oscillations on the scale of D1 , respectively. This description becomes invalid when tunneling matrix elements are comparable in magnitude with the matrix elements of YY 0 coupling. As a result, tunneling and
intrawell transitions are no longer separated in time. Nevertheless, the PIA remains valid in this case.
The propagator, Eq. (28), can also be used to study incoherent processes, if the interactions responsible
for the dissipation are specied. This problem is the cornerstone of the theory of radiationless processes in
general and of incoherent tunneling dynamics in particular. Dissipative tunneling arises from system-bath
interactions and manifests itself in a wide variety of non-rigid molecules in solids. However, the details of
the system-bath interactions are almost never specied on a microscopic level. Instead of a set of coupling
constants between internal and bath degrees of freedom, phenomenological parameters like spectral densities or friction coecients are introduced. Even this simplied problem is solved only for 1D tunneling
system interacting with a harmonic bath. On the other hand, molecular force elds are now easily computed
using the dierent methods of molecular dynamics. The results can be represented in the form of time
dependent atomic displacements, which lead to a 3N 6 dimensional Hamiltonian with time dependent
frequencies and coupling constants. Through a Fourier transformation, these uctuations of parameters
can be represented in the form of an individual bath for each of internal degree of freedom. These uctuations destroy the coherence near resonances, and this eect gives rise to dissipative tunneling. An additional channel of vibrational relaxation arises from tunneling induced transitions between dierent
vibrational states. Even though much additional work is required in this direction, the results of this paper
open the way to a study of dissipative tunneling on a microscopic level.
Appendix A. Second-order action along the extreme tunneling trajectory
The equations of instanton motion in the upside-down PES, Eq. (1), are:
oV
;
X
oX

oV
Yk
oYk

A:1

416

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

For each L, G, or Sq transverse coordinate, the second equation is split into equations for displacements
proportional to Ck and akk0 Ck0 , when terms of order a2kk0 Ck0 are neglected:
X
Y k yk
ykk0
k0

yk x2k yk

Ck X ;

kL
2

ykk 0 x2k ykk0

Ck 1 X ; k  G
8


>
Ck 0
2
>
0
0
>
X
y

X
;
a
k
< kk
x2

k; k 0  L

k0

A:2

k  L; k 0  G; Sq

akk0 Xyk0 ;
>
>
>
:
akk0 X 2 yk0 ;

k; k 0  G; Sq

Since for Sq vibrations yk 0, and as YY 0 couplings between Sq and L or G vibrations are forbidden by
symmetry, the ETT does not deviate from 1D path along Sq modes. Since the variables yk aect only the
second-order equation for the X-motion [11], in the rst-order equations (A.2), X t can be replaced by the
solution of the zero-order equation for the uncoupled X-motion:
oV0
X
oX

A:3

Eq. (A.2) is integrated in parallel and give:


Z 1



Ck
exp xk t t0 f t0 dt0
yk
2xk 1

A:4

where f t X t and f t 1 X 2 t for L and G vibrations, respectively. Expanding the integral in


Eq. (A.4) into a series, one readily derives the expression for the ETT in the high-frequency limit. For L
vibrations, transverse displacements have the following form:
yk

Ck
Ck
X 4 X ;
x2k
xk

ykk0

akk0 Ck0 2 
X X
x4k0

An action between the minima is given by:


Z 1
H X ; PX ; fYk ; Pk g dt
W
1

A:5

A:6

where the coordinates and conjugate momenta obey Eq. (A.1). Using Eqs. (A.2) and (A.3) we rewrite Eq.
(A.6) in the following form:
X
X
dWk
dWkk 0
A:7
W  W0
k

where
W0

kk 0


1 2
PX V0 X dt
2

A:8

represents the 1D action, while dWk describe corrections due to L and G couplings:
(
dWk


R1 
C2k 1 X Yk xCk2 X dt;
k
R1
Ck
1 X 2 Yk dt;
1
2

kL
kG

A:9

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

417

dWk 0 for Sq vibrations. Using Eq. (A.4) and integrating Eq. (A.9) by parts, we specify the functions in
Eq. (11):
Z



xk 1
L
exp xk t t0 X_ tX_ t0 dt dt0
F xk
4 1
Z
A:10





xk 1
F G xk
exp xk t t0 1 X 2 t 1 X 2 t0 dt dt0
4 1
where X t is given by Eq. (A.3). The rst of two Eqs. (A.10) was derived in Ref. [8] using the Feynman
2
Vernon formalism of inuence functional [78]. For the quartic potential, V0 1 X 2 , functions F L and
G
2
F dier only by the factor x [53]. The third term in Eq. (A.7) describes the eect of Sq and YY 0
couplings. Its expression is given in Appendix A to Ref. [17].
Appendix B. Squeezed and kinematic coupling contributions to HamiltonJacobi equation
The HJE for a system with a X-dependent kinematic matrix is derived directly from the time-independent Schr
odinger equation when kinetic energy operator is given by Eq. (12), (see for example Ref. [11]):
1 X oW oW
Gij
V Q
B:1
2 ij
oQi oQj
where Q X ; fYk g. The HJE, Eq. (B.1), of order c involves no derivatives of the kinematic matrix, which
appears in the TE at order c0 (see Appendix C). The X-dependent kinematic matrix is dened by Eq. (6).
The zero-order action is chosen as a solution of the HJE with the unit kinematic matrix. The rst-order
HJE is additive with respect to diagonal Sq couplings and non-diagonal L, G, and YY 0 couplings:
W1 W11 W12 W13

B:2

which satisfy the following equations:


2

oW0 oW11 X
oW11
1
1 X
2 oW0

xk Yk
gXX X

akk x2k gkk X 2 Yk2


oX oX
oY
oX
2
2
k
k
k

B:3



oW0 oW12 X
oW12 X
oW0

Yk
xk Yk

Ck fk X xk gXk f~k X
oX oX
oYk
oX
k
k

B:4

where fk X X , f~k X X 2 and fk X 1 X 2 , f~k X X for L and G vibrations, respectively. The


eect of YY 0 coupling is described by the following equation:


oW0 oW13 X
oW13 X
1

xk Yk

akk0 xk xk0 gkk0 fkk0 X Yk Yk0


B:5
oX o X
oYk
2
k
kk 0
where fkk0 X X , k  L, k 0  G and fkk0 X X 2 otherwise. Eqs. (B.4)(B.6) show that kinematic couplings renormalize the potential L, G, and Sq couplings in the rst-order HJE. The variation of eective
mass of the longitudinal motion is given by the rst right-hand side term in Eq. (B.4). The linearized
equations (B.4)(B.6) are readily solved using the characteristic trajectories from the zero-order HJE. The
general considerations are given in Refs. [11,13]. For brevity, we write down these trajectories for one
branch of the solution of the HJE corresponding to the localized wave function of the left well, X0 1,
when the instanton starts at t ! 1:

418

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

oW0
;
X_
oX

Y_ k xk Yk

B:6

Along these characteristic trajectories, the rst-order action is reduced to quadratures:


Z X
X 1
1
oW0
dx
x2
Fkk Yk2
W11 gXX
2
o
x
1
k
W12

X 1
Fk Yk ;

W13

where
1

X
k

B:7

Fkk0 Yk Yk0

B:8

Z
 t


1
akk x2k gkk
exp 2xk t t0 X 2 t0 dt0
2
1


Z t


oW0
0
0
0
~

exp xk t t Ck fk X t xk gXk fk X t
dt0
o X tt0
1

Z t


1
akk0 xk x k0 gkk0
exp xk xk0 t t0 fkk0 X t0 dt0
2
1

Fkk
1

Fk

Fkk0

B:9

where X t is given by Eq. (A.3). Eqs. (B.7)(B.9) generalize the relations for a quartic potential, derived in
Ref. [13], to the case of bistable 1D potentials of arbitrary shape. The second-order action W2 satises a
partial dierential equation with the same characteristic trajectories, Eq. (B.6), and with the right-hand side
involving the combinations of kinematic and potential coupling terms:

2 X(

2
oW0 oW2 X
oW2
1 oW1
1 oW1
2 oW0 oW1
2 oW0 oW1

xk Yk
gXX X

gkk X

oX oX
oYk
oX oX
oYk oYk
2 oX
2 oYk
k
k

)
1
oW0 oW1 oW1 oW0
gXk f~k X

oX oYk
oX oYk
2


1X
oW0 oW1 oW1 oW0
gkk 0 fkk0 X

B:10
2 kk0
oYk oYk0 oYk oYk0
Eq. (B.10) denes a second-order action as quadratic form of {Yk } with X-dependent coecients [11,13].
Z X
X 2
X 2
X 2
oW0
2
dx
x4
dWk X
Fk X Yk
Fkk0 X Yk Yk0
B:11
W2 gXX
ox
1
k
k
kk 0
where
2

dWk

2
Fk

2
Fkk0

1
2

Z

Z

2
1
Fk t0 dt0 gXk

XZ

k0



oW0
0
f~k X t0
dt
oX tt0
1
t

B:12

 1
1
exp xk0 t t0 Fk0 t0 Fkk 0 t0 dt0


1
1

dF
dF
0
k
k
exp xk xk0 t t0

dX dX

dt 2
tt0

X
k 00

!
1
1
Fkk00 t0 Fk0 k00 t0

B:13

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

419

At X 1, t ! 1, the rst sum in Eq. (B.11) determines contributions of L and G vibrations to an action
between the minima. The eect of YY 0 couplings arises in the higher order HJE. The second and third sums
in Eq. (B.11) describe the departure of the characteristic trajectory, Eq. (B.6), from the 1D-path and the
X-dependence of the transverse frequencies, respectively. As shown in Ref. [11], the characteristic trajecL
R
tories for the left and right wells,fY k g and fY k g, dened by following equations:

o  LR
LR
LR
W0
W2
W1
0
B:14
oYk
do not coincide with the ETT. The equations of instanton motion are satised by the half sum of the curves,
Eq. (B.14), starting from the left X0 1 and right X0 1 wells:

1 L
R
Yk
Yk Yk
B:15
2
In order to nd the action for the wave function of the right well, the second branch of the HJE,
Y_ k xk Yk , is used. For this branch, the momenta conjugate to Yk change sign, so that the kinematic
coupling coecients gXk are must be changed to gXk in order to provide the invariance of the Hamitonian
with respect to time inversion. The potential and kinematic terms in Eq. (B.4) are dierent with respect to
inversion of X and the change gXk ! gXk , indicated above, provides the correct symmetry of the ETT with
respect to the dividing surface. Thus, the symmetrized combination of HJE solutions corresponds to the
ETT, while the set of second derivatives with respect to Y determines the width of the reactive channel
centered around the ETT. Although these characteristics are similar to those of the RPHF, an important
dierence exists between the MEP and the ETT. As readily seen from Eqs. (A.9) and (B.7)(B.9), an action
along the ETT results from the integration over all previous instants in time, when the system evolves from
the initial state to the given point of phase space. This delay is ignored in the RPHF.
The solution of the third- and fourth-order HJE is performed in the same way and leads to an expansion
of an action in power series of Yk . X-dependent coecients of this expansion are given in Appendix B to
Ref. [17].
Appendix C. Transport equation
The TE has the following form:




1X
oW oA oW oA
o2 W
1 oGij oW 1 oGij oW
Gij

A Gij

EA
2 ij
oQi oQj oQj oQi
oQi oQj 2 oQi oQj 2 oQj oQi

C:1

where E stands for the set of eigenvalues characterized by vibrational quantum numbers n; fnk g. The zeroorder eigenvalues correspond to a separable system:


 X 



gXX 
1
gkk 
akk
1
0
xk 1
n
1 2

C:2
nk
Enfnk g x 0 1
2
2
xk
2
2
k
where x0 relates to the eigenfrequency of the longitudinal vibration with unit mass and where diagonal
potential and kinematic Sq couplings (kinematic couplings produce variations of eective masses associated
with X and Yk motions) are taken into account. The couplings aect only the second-order eigenvalues, so
0
that the rst-order TE involves Enfnk g and the additive contributions of potential and kinematic couplings.
As shown in Ref. [11], Sq couplings renormalize the zero-order prefactors:

420

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422


0
0
A~nfnk g Anfnk g 1 Gnfnk g X

C:3

where the factors Gnfnk g obey the following equation:




!

 
oW0 oG X
oG 1
1
oW0 X
akk
1
2
2

x0 gXX X n
xk Yk

xk gkk 2 X nk
gXX X
oX oX
oX
xk
oYk 2
2
2
k
k

C:4

The right-hand side of Eq. (C.4) is independent of Yk and the factors Gnfnk g are readily calculated by integration of the right-hand side over X t. The auxiliary functions, Eq. (25), are eigenfunctions of the
transformed Hamiltonian:
c

^
2
e
C:5
H
exp cW1 H^ exp cW1 H^ R^ O rW1
2
where
R^ rW1 r 12DW1

C:6

If the non-separable Hamiltonian is divided in the separable part H^0 and the perturbation V^int summing all
coupling terms, the initial and transformed Hamiltonians, Eq. (C.5), have the same set of the rst-order
eigenvalues and the latter involve the new perturbation operator:
V^1 V^int R^

C:7

Taking into account that the rst-order action W1 obeys an equation, which is the sum of Eqs. (B.3)
(B.5):
rW0 rW1 V^int

C:8

one readily obtains:


0
0
V^1 Wnfnk g exp cW0 R^ A~nfnk g

C:9

where terms of order c1 of the operators V^int and R^ cancel each other. The rst-order coecients in expansion, Eq. (22), have the following form:

D
E
0
0
Wn0 fn0n g V^1 Wnfnn g
1
C:10
anfnn gn0 fn0n g
0
0
Enfnk g En0 fn0 g
k

The second-order prefactors are constructed using the operator V^2 of order c0 , given by the expression:
2
V^2 crW0 rW2 rW2 r 12DW2 Enfnk g

C:11

where W2 stands for the second-order action dened by Eq. (B.10), the rst term eliminates the rest term in
Eq. (C.5), and the last term is the second-order correction to the eigenvalues. The second-order coecients
in expansion, Eq. (22), consist of the matrix elements of both operators (C.7) and (C.11):

anfnk gn0 fn0 g


k


D
E
0
0
Wn0 fn0n g V^2 Wnfnn g
0
Enfnk g

0
En0 fn0 g
k

X
n00 fn00k g





0
0
0
0
Wn00 fn00 g V^1 Wnfnk g
Wn0 fn0 g V^1 Wn00 fn00 g
k

0
Enfnk g

0
En00 fn00 g
k

0
En00 fn00 g
k

0
En0 fn0 g
k

C:12

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422

421

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]

W.F. Rowe, R.W. Duerst, E.B. Wilson, J. Am. Chem. Soc. 98 (1976) 420.
S.L. Baughcum, R.W. Duerst, W.F. Rowe, Z. Smith, E.B. Wilson, J. Am. Chem. Soc. 103 (1981) 6296.
S.L. Baughcum, Z. Smith, E.B. Wilson, R.W. Duerst, J. Am. Chem. Soc. 106 (1984) 2260.
P. Turner, S.L. Baughcum, S.L. Coy, Z. Smith, J. Am. Chem. Soc. 106 (1984) 2265.
D.W. Firth, K. Beyer, M.A. Dvorak, S.W. Reeve, A. Grushow, K. Leopold, J. Chem. Phys. 94 (1991) 1812.
T. Baba, T. Tanaka, I. Morino, K.M.T. Yamada, K. Tanaka, J. Chem. Phys. 110 (1999) 4131.
J.R. de la Vega, Acc. Chem. Res. 15 (1982) 185.
V.A. Benderskii, D.E. Makarov, C.A. Wight, Chemical Dynamics at Low Temperatures, Wiley-Interscience, New York, 1994.
H.P. Trommsdor, Adv. Photochem. 24 (1998) 147.
V.A. Benderskii, S.Yu. Grebenshchikov, E.V. Vetoshkin, L. von Laue, H.P. Trommsdor, Chem. Phys. 219 (1997) 119.
V.A. Benderskii, E.V. Vetoshkin, L. von Laue, H.P. Trommsdor, Chem. Phys. 219 (1997) 143.
V.A. Benderskii, E.V. Vetoshkin, H.P. Trommsdor, Chem. Phys. 234 (1998) 153.
V.A. Benderskii, E.V. Vetoshkin, Chem. Phys. 234 (1998) 173.
V.A. Benderskii, E.V. Vetoshkin, H.P. Trommsdor, Chem. Phys. 244 (1999) 273.
V.A. Benderskii, E.V. Vetoshkin, H.P. Trommsdor, Chem. Phys. 244 (1999) 299.
V.A. Benderskii, E.V. Vetoshkin, Chem. Phys. 257 (2000) 203.
V.A. Benderskii, E.V. Vetoshkin, H.P. Trommsdor, Chem. Phys. 262 (2000) 369.
K. Fukui, Pure Appl. Chem. 54 (1982) 1825.
S. Kato, H. Kato, K. Fukui, J. Am. Chem. Soc. 99 (1977) 684.
T. Carrington, W.H. Miller, J. Chem. Phys. 84 (1986) 4364.
W.H. Miller, N.C. Handy, J.E. Adams, J. Chem. Phys. 72 (1980) 99.
W.H. Miller, J. Phys. Chem. 87 (1983) 3811.
D.G. Truhlar, B.C. Garrett, Ann. Rev. Phys. Chem. 35 (1984) 159.
J. Bicerano, H.F. Schaefer, W.H. Miller, J. Am. Chem. Soc. 105 (1983) 2550.
V.A. Benderskii, V.I. Goldanski, D.E. Makarov, Chem. Phys. 159 (1992) 29.
V.K. Babamov, R.A. Marcus, J. Chem. Phys. 74 (1978) 1790.
M.Ya. Ovchinnikova, Chem. Phys. 36 (1979) 85.
V.A. Benderskii, V.I. Goldanskii, A.A. Ovchinnikov, Chem. Phys. Lett. 73 (1980) 492.
N. Makri, W.H. Miller, J. Chem. Phys. 91 (1989) 7026.
N. Makri, W.H. Miller, J. Chem. Phys. 86 (1987) 1451.
W.H. Miller, B.A. Ruf, Y.-T. Chang, J. Chem. Phys. 89 (1988) 6298.
N. Shida, P. Barbara, J.E. Almlof, J. Chem. Phys. 91 (1989) 4061.
T.D. Sewell, Y. Guo, D.L. Thompson, J. Chem. Phys. 103 (1995) 8557.
B.C. Garrett, D.G. Truhlar, A.F. Wagner, T.H. Dunning, J. Chem. Phys. 78 (1883) 4400.
E. Bosh, H. Moreno, J.M. Lluch, J. Bertran, J. Chem. Phys. 93 (1990) 5685.
V.A. Benderskii, P.G. Grinevich, D.E. Makarov, Chem. Phys. 170 (1993) 275.
V.A. Benderskii, S.Yu. Grebenshchikov, D.E. Makarov, E.V. Vetoshkin, Chem. Phys. 185 (1994) 101.
V.P. Maslov, M.V. Fedoriuk, Semiclassical Approximation in Quantum Mechanics, Reidel, Boston, 1981.
J.B. Delos, Adv. Chem. Phys. 65 (1986) 161.
S. Takada, H. Nakamura, J. Chem. Phys. 100 (1994) 98.
S. Takada, H. Nakamura, J. Chem. Phys. 102 (1995) 3977.
G.V. Milnikov, A.J.C. Varandas, J. Chem. Phys. 111 (1999) 8302.
A.M. Polyakov, Nucl. Phys. B. 120 (1977) 429.
S. Coleman, Aspects of Symmetry, Cambridge University Press, Cambridge, 1985.
R. Rajaraman, Solitons and Instantons, North-Holland, Amsterdam, 1982.
V.A. Benderskii, V.I. Goldanskii, D.E. Makarov, Phys. Reps. 233 (1993) 195.
V.A. Benderskii, S.Yu. Grebenshchikov, G.V. Milnikov, E.V. Vetoshkin, Chem. Phys. 188 (1994) 19.
V.A. Benderskii, S.Yu. Grebenshchikov, G.V. Milnikov, Chem. Phys. 194 (1995) 281.
V.A. Benderskii, S.Yu. Grebenshchikov, G.V. Milnikov, Chem. Phys. 198 (1995) 19.
Z. Smedarchina, W. Siebrand, M.Z. Zgierski, J. Chem. Phys. 103 (1995) 5326.
Z. Smedarchina, W. Siebrand, M.Z. Zgierski, J. Chem. Phys. 104 (1995) 1203.
Z. Smedarchina, W. Caminatti, F. Zerbetto, Chem. Phys. Lett. 237 (1995) 279.
Z. Smedarchina, A. Fernandez-Ramos, M.A. Rios, J. Chem. Phys. 106 (1997) 3856.
A. Fernandez-Ramos, Z. Smedarchina, M.Z. Zgierski, W. Siebrand, D O T T 1 . 0 , a computer program to calculate hydrogentunneling splittings in molecular spectra, National Research Council of Canada, Ottawa, KIA, OR6.

422
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]
[65]
[66]
[67]
[68]
[69]
[70]
[71]
[72]
[73]
[74]
[75]
[76]
[77]
[78]

V.A. Benderskii et al. / Chemical Physics 262 (2000) 393422


Z. Smedarchina, M.Z. Zgierski, W. Siebrand, P.M. Kozlowski, J. Chem. Phys. 109 (1998) 1014.
V. Barone, C. Adamo, J. Chem. Phys. 105 (1996) 11007.
S. Sadhukhan, D. Munoz, C. Adamo, G.E. Scuseria, Chem. Phys. Lett. 306 (1999) 83.
R. Meyer, H.H. G
unthard, J. Chem. Phys. 50 (1969) 353.
M.J. Frish, A.C. Scheiner, H.F. Schaefer, J. Chem. Phys. 82 (1985) 4194.
J.S. Binkley, M.J. Frish, H.F. Schaefer, Chem. Phys. Lett. 126 (1986) 1.
Z. Latajka, S. Scheiner, J. Phys. Chem. 96 (1992) 9764.
K.B. Wiberg, C.M. Hadad, T.J. LePage, C.M. Breneman, M.J. Frish, J. Phys. Chem. 96 (1992) 671.
S. Scheiner, J. Mol. Struct. (Theochem) 307 (1994) 65.
K. Luth, S. Scheiner, J. Phys. Chem. 98 (1994) 3582.
C.W. Bauschlicher, H.J. Partridge, J. Chem. Phys. 103 (1995) 1768.
S. Sirois, E.I. Proynov, D.T. Nguyen, D.R. Salahub, J. Chem. Phys. 107 (1997) 6770.
R.S. Brown, A. Tse, T. Nakashima, R.C. Hadson, J. Am. Chem. Soc. 101 (1979) 3157.
R.V. Stanton, K.M. Merz, J. Chem. Phys. 101 (1994) 6658.
C. Lee, W. Yang, R.G. Parr, Phys. Rev. B. 37 (1988) 785.
A.D. Becke, J. Chem. Phys. 98 (1995) 5648.
T. Chiavassa, P. Roubin, L. Pizzala, P. Verlaque, A. Allouche, J. Phys. Chem. 96 (1992) 10659.
A. Schmid, Ann. Phys. (NY) 170 (1986) 333.
R.G. Littlejohn, J. Stat. Phys. 68 (1992) 7.
S. Keshavamurthy, W.H. Miller, Chem. Phys. Lett. 218 (1994) 189.
B.W. Spath, W.H. Miller, Chem. Phys. Lett. 262 (1996) 486.
K. Kay, J. Chem. Phys. 100 (1994) 4377.
K. Kay, J. Chem. Phys. 107 (1997) 233.
R.P. Feynman, F.I. Vernon, Ann. Phys. (NY) 24 (1963) 118.

Вам также может понравиться